Transverse mixing coefficient (TMC) is known as one of the most effective parameters in the two-dimensional simulation of water pollution, and increasing the accuracy of estimating this coefficient will improve the modeling process. In the present study, genetic algorithm (GA)-based support vector machine (SVM) was used to estimate TMC in streams. There are three principal parameters in SVM which need to be adjusted during the estimating procedure. GA helps SVM and optimizes these three parameters automatically in the best way. The accuracy of the SVM and GA-SVM algorithms along with previous models were discussed in TMC estimation by using a wide range of hydraulic and geometrical data from field and laboratory experiments. According to statistical analysis, the performance of the mentioned models in both straight and meandering streams was more accurate than the regression-based models. Sensitivity analysis showed that the accuracy of the GA-SVM algorithm in TMC estimation significantly correlated with the number of input parameters. Eliminating the uncorrelated parameters and reducing the number of input parameters will reduce the complexity of the problem and improve the TMC estimation by GA-SVM.

  • Genetic algorithm (GA)-based support vector machine (SVM) was used to estimate TMC in streams.

  • Sensitivity analysis showed that the accuracy of GA-SVM algorithm in TMC estimation significantly correlated with the number of input parameters.

Increasing the accuracy of modeling the process of pollution release into streams will increase the ability to control the quality of streams and thereby reduce environmental damage. Therefore, the capability to estimate the transport of pollutants in streams and waterways has always been a considerable issue in many industrial and environmental projects (Abderrezzak et al. 2015). After being discharged into a river, contaminants and effluents mix with water of the river being transported to the downstream (Seo & Cheong 1998). The effluent is spread vertically, transversely, and longitudinally by advective and dispersive transport processes. In a shallow stream, after contamination is rapidly mixed throughout the depth, the transmission will occur in the longitudinal and transverse directions (Ahmad et al. 2011). A full cross-sectional mix will not be achieved, unless the pollutant travels the long distances which are generally not within the length of practical interest (Beltaos 1980). The length required for full cross-sectional mixing of contaminations is approximately 20 and 200 times the upper width for a rough and a smooth flow, respectively (Fischer 1967). Transverse mixing plays an important role in determining the effect of contaminants under steady-state conditions. This parameter has an important effect in water quality management; especially in a case of point source discharges or tributary inflows (Rutherford 1994; Boxall & Guymer 2003). According to Figure 1, for the effluent mixing process in rivers, three stages are considered: (1) mixing near to the discharging point due to initial momentum and flow buoyancy (between A and B zones); (2) transverse mixing due to turbulence (secondary turbulence transfer) and its secondary flows (between B and C zones); and (3) dispersion due to longitudinal shear flow (after C zone) (Fischer et al. 1979).

Figure 1

General steps of pollution dispersion in a stream (Fischer et al. 1979).

Figure 1

General steps of pollution dispersion in a stream (Fischer et al. 1979).

Close modal
The distribution of tracer concentration can be written in a two-dimensional model according to the principle of mass conservation (Rutherford 1994; Sharma & Ahmad 2014):
formula
(1)
where t is the time; H is depth of flow ; C is the depth-averaged tracer concentration ; and x are the transverse and longitudinal directions, respectively; and are the velocities in the z and x directions , respectively; and are the depth-averaged dispersion coefficients in transverse and longitudinal directions . By assuming that longitudinal dispersion of tracer has not begun yet for the uniformly flowing stream, the time differentiation of Equation (1) will be zero (Sharma & Ahmad 2014). Also, by assuming a uniform flow and , Equation (1) can be simplified to:
formula
(2)

The above equation has been used in many studies (Krishnappan & Lau 1977; Lau & Krishnappan 1981; Demetracopoulos 1994; Ahmad 2008; Aghababaei et al. 2017; Huai et al. 2018; Zahiri & Nezaratian 2020). More investigations on the role of the effective parameters in transverse mixing would be required due to the complexity of the transverse mixing mechanism (Aghababaei et al. 2017). Thus, predicting the transverse mixing coefficient (TMC) for known flow conditions in a stream for accounting the pollutant concentration at any location downstream of the injection site is genuinely essential (Azamathulla & Ahmad 2012). Generally, there are three approaches for predicting the TMC in stream mixing. Empirical methods have developed equations using the hydraulic and geometric dataset of rivers and experimental studies in order to establish a relationship for and theoretical methods have used the concept of shear flow to derive the dispersion coefficient (Baek & Seo 2013). Moreover, many researchers have recently used powerful predictive tools to find solutions for complex engineering problems. The significance of dispersion coefficients in water quality modeling and the complexity of the pollutant emission and mixing process have considerably increased the importance of using these tools (Zahiri & Nezaratian 2020). Soft computing techniques such as fuzzy-neural inference system-based principal component analysis (ANFIS-based PCA), particle swarm optimization method (PSO), artificial neural network (ANN), genetic expression programming (GEP), differential evolution (DE), decision tree (M5), support vector machine (SVM), and fuzzy-neural inference system (ANFIS) have been widely used to estimate longitudinal dispersion coefficient in streams by Parsaei et al. (2018), Alizadeh et al. (2017), Antonopoulos et al. (2015), Sattar & Gharabaghi (2015), Li et al. (2013), Etemad-Shahidi & Taghipour (2012), Azamathulla & Wu (2011) and Riahi-Madvar et al. (2009). Azamathulla & Ghani (2011), Azamathulla & Ahmad (2012), Aghababaei et al. (2017), and Zahiri & Nezaratian (2020), tried to predict the TMC accurately by using decision tree (M5), multivariate adaptive regression splines (MARS), particle swarm optimization method (PSO), multiple linear regression (MLR), genetic algorithm (GA), genetic programming for symbolic regression (GPSR), and GEP. Soft computing techniques used by the above-mentioned researchers have less statistical errors and higher accuracy than empirical methods in TMC prediction (Zahiri & Nezaratian 2020). According to previous studies, there is a strong relationship between the TMC and channel parameters such as channel width, flow depth, shear velocity, friction factor, curvature and sinuosity (Fischer 1967; Beltaos 1979; Lau & Krishnappan 1981; Stefanovic & Stefan 2001; Boxall & Guymer 2003). Table 1 shows some of the most well-known equations proposed for calculating the TMC.

Table 1

Some of the empirical and data-driven models for estimation of TMC

is the TMC (m2/s), H is the flow depth (m), is a bed shear velocity (m/s), W is a channel width (m), is sinuosity coefficient and is a Froude number.

Each of these mentioned algorithms has its strengths and weaknesses that may not be able to predict complex phenomena such as TMC accurately. Selecting several meta-heuristic algorithms correctly and using them simultaneously will increase accuracy and decrease errors in target values’ estimation. Selecting an algorithm as the main algorithm along with an auxiliary algorithm that can improve the weaknesses of the main algorithm will lead to developing a hybrid algorithm with higher performance. In previous investigations, several hybrid algorithms were used to estimate some of the complex phenomena and, consequently, the ability of these algorithms was proven completely (Pourbasheer et al. 2009; Wang et al. 2013; Li & Kong 2014; Zhou et al. 2016). In this study, two common algorithms were used to develop a hybrid algorithm: support vector machine (SVM) as the main algorithm and genetic algorithm (GA) as the auxiliary algorithm. Connecting GA to SVM allows us to estimate optimal values of SVM's adjustable parameters in the shortest time and increase predicting accuracy. The purpose of this study is developing an SVM-GA algorithm by using 232 published datasets and making a comparison of its performance with previous models. In addition, sensitivity analysis has been performed on the developed model to determine the effect of input parameters in the TMC modeling.

Data

In the present study, 232 data points (see Supplementary material) were collected from the technical literature (Yotsukura et al. 1970; Holley & Abraham 1973; Krishnappan & Lau 1977; Beltaos 1979; Rutherford 1994; Jeon et al. 2007; Baek & Seo 2008; Lee & Seo 2013). It must be added that 183 and 49 dataset have been collected from straight and meandering streams, respectively. In addition, the dataset contains geometrical and hydraulic characteristics, including channel width, channel depth, average velocity, shear velocity, Froude number, sinuosity, and TMC. Sinuosity was used to demonstrate horizontal irregularities in meandering streams (Aghababaei et al. 2017). Table 2 illustrates a statistical analysis of all variables.

Table 2

Descriptive statistics for the TMC database

ParameterWHUU*W/HU/U*FrSnε z/HU*ε z
Min 0.200 0.013 0.040 0.005 1.670 2.051 0.018 1.000 0.054 0.000034 
Max 320.000 5.250 1.750 0.163 287.500 28.571 0.971 3.330 2.400 0.215 
Avg 15.950 0.304 0.308 0.026 26.710 12.976 0.285 1.108 0.238 0.007 
SD 51.237 0.709 0.271 0.023 34.995 5.447 0.181 0.371 0.249 0.025 
Skewness 4.246 4.506 2.947 2.379 3.797 0.196 0.866 4.974 4.510 5.246 
ParameterWHUU*W/HU/U*FrSnε z/HU*ε z
Min 0.200 0.013 0.040 0.005 1.670 2.051 0.018 1.000 0.054 0.000034 
Max 320.000 5.250 1.750 0.163 287.500 28.571 0.971 3.330 2.400 0.215 
Avg 15.950 0.304 0.308 0.026 26.710 12.976 0.285 1.108 0.238 0.007 
SD 51.237 0.709 0.271 0.023 34.995 5.447 0.181 0.371 0.249 0.025 
Skewness 4.246 4.506 2.947 2.379 3.797 0.196 0.866 4.974 4.510 5.246 
Table 2 implies that the studied cases varied from narrow rivers () to very wide rivers (). , which is known as friction term and represents the hydrodynamic and roughness of the canal bed (Seo & Cheong 1998), varied from 0.026 to 28.571. This range of variations indicates the usage of a wide range of streams with various geometrical and hydraulic features in this study, the results of which can be related to many streams with different characteristics. The dataset was randomly divided into two sets, training (75% of the data) and testing (25% of the data). Although many unknown parameters may affect the TMC, according to previous studies, the key parameters affecting the mixing process during steady flow in natural streams can be stated as follows:
formula
(3)
where is the fluid density; is fluid viscosity; and are bed shape factor and sinuosity, respectively; and g is gravity. Fischer et al. (1979) and Jeon et al. (2007) expressed the relation below in terms of dimensionless parameters by using Buckingham Pi theorem:
formula
(4)
where is the friction term; is the channel width to flow depth ratio; is Froude number; and is Reynolds number. Bed shape factor, , and sinuosity, , indicate vertical and transverse irregularities in natural streams, respectively (Etemad-Shahidi & Taghipour 2012). By developing secondary currents and shear flow, transverse and vertical irregularities affect the mixing processes in streams (Seo & Cheong 1998). Generally, the flow in natural streams is usually fully turbulent, so Reynolds number could be eliminated from Equation (4) as a first approximation (Seo & Cheong 1998; Kashefipour & Falconer 2002). Bed shape factor could also be eliminated from this equation as Froude number and dimensionless roughness factor can reflect the other effects of bed material roughness and bed slope (Sattar & Gharabaghi 2015). Finally, the best dimensionless form of based on previous findings such as those of Yotsukura & Sayre (1976), Deng et al. (2001), Jeon et al. (2007), Azamathulla & Ahmad (2012), Aghababaei et al. (2017), and Zahiri & Nezaratian (2020) can be written as follows:
formula
(5)
where represents the dimensionless parameter of and it will be used as the target parameter in this research. The correlations between all input and output parameters are displayed in Figure 2.
Figure 2

Correlations between all input and output parameters.

Figure 2

Correlations between all input and output parameters.

Close modal

Based on Figure 2, there is no considerable correlation between the input variables, thus the problems that could arise in analysis from exaggerating the strength of the relations between variables, would be eliminated (Sattar & Gharabaghi 2015). It should be noted that the average of each parameter in training and testing subsets is equal to (13.36, 25.63, 0.29, 1.12, 0.25) and (11.91, 30.18, 0.27, 1.06, 0.20), respectively.

Support vector machine (SVM)

Vapnik (1995) proposed a nonlinear regression predicting method called support vector machine (SVM) which was usable to solve pattern recognition, highly nonlinear classification and regression problems. Maximizing the accuracy of prediction or minimizing the difference between the outputs and targets was the purpose of developing the SVM (Parsaie & Haghiabi 2017a, 2017b; Parsaie et al. 2019). For this purpose, the input parameters are mapped into a high-dimensional linear feature space by a nonlinear transformation to construct the optimal decision function. The dot product operation in the higher dimensional feature space is replaced by the kernel function in the original space, and by the finite sample training, the global optimal solution is obtained (Zhou et al. 2016). In the current study, SVM is used for predicting the TMC as the main algorithm, which is briefly described below.

If data is assumed as training set, where is the input vector,, is the output, and n is the number of data pairs, the regression function of SVM which is called SVR will be formulated as follows:
formula
(6)
where represents the transposed form of vector; b is a bias; and can be obtained through some restricted rules. This function can describe the observed output y with an error tolerance . would be considered as a nonlinear transfer function mapping the input vectors into a high-dimensional feature space which, theoretically, even a simple linear regression will be able to overcome the complexity of nonlinear regression of the input space (He et al. 2014). The tolerated errors within the extent of the -tube, as well as the penalized losses when data concern the outside of the tube, are defined by Vapnik's -insensitive loss function as:
formula
(7)
After that, the SVM problem can be formulated as the optimization problem as below:
formula
(8)
formula
(9)
where the constant C is called a penalty factor and shows the penalty degree of the sample with error exceeding (Liu & Jiao 2011). Here, the value of C is set to 1 which shows the complexity of the model is as important as the empirical error. Also, and are introduced as slack variables that specify the upper and lower errors of training subject to the error tolerance . These variables express the distance difference between actual values and the corresponding boundary values of -tube. Figure 3 depicts the mentioned situation graphically. SVM reduces under-fitting and over-fitting problems by minimizing and which are called the regularization and training error terms, respectively.
Figure 3

Nonlinear SVM with Vapnik's e-insensitive loss function.

Figure 3

Nonlinear SVM with Vapnik's e-insensitive loss function.

Close modal
Thus, the dual Lagrangian form will be yielded as follows by considering Lagrangian multipliers and Karush–Kuhn–Tucher condition in Equation (9):
formula
(10)
formula
(11)
where and are Lagrangian multipliers that satisfy equalities; , and also, represents the Lagrange function. The Lagrange multiplier terms related to the data accumulating the inside of the -insensitive tube will be considered to be zero. The final regression function is calculated only by using the datasets with non-zero coefficients which are known as the support vectors. There are two groups of support vectors: margin support vectors and error support vectors (Noori et al. 2011). In the first group, the support vectors have absolute values of the weights less than C and in the second group, equal to C. In other words, the support vectors, which are located outside and on the margin of the insensitive tube, are called the error support vectors and the margin support vectors, respectively (Figure 3). For changing the dimensionality of the input space to perform the regression or classification task with more confidence, kernel functions are used (Azamathulla & Wu 2011). These functions yield the inner products in the feature space and . A kernel function plays the most significant role to simplify the learning process by changing the representation of the data in the feature space. Thus, although the data may be non-separable in the original input space, an appropriate choice of a kernel function allows the data to be highly separable in the feature space (Patil et al. 2012). If there is no prior knowledge about data features, radial basis function (RBF) will be recommended as one of the most popular kernel functions which is being used in different scientific fields (Roushangar & Koosheh 2015). For this reason, in this study, RBF was used as the kernel function of the SVM model for the TMC prediction.
formula
(12)
where is a kernel function and is the parameter of the RBF kernel function.

Genetic algorithm (GA)

According to the mechanisms of genetics and Darwin's natural selection principles, John Holland in 1975, proposed a heuristic search method and called it the genetic algorithm (GA). This method was named after biological processes of inheritance, mutation, natural selection, and the genetic crossover that happens when parents mate to produce offspring (Goldberg 1989). Technically, there are four differences between the structure of GA and other traditional optimization algorithms (Goldberg 1989):

  • The GA typically uses a coding of the decision variable set instead of decision variable itself.

  • The GA searches from a population of decision variable sets instead of a single decision variable set.

  • The GA uses the objective function itself instead of the derivative information.

  • The GA algorithm uses probabilistic instead of deterministic, search rules.

In the last decade, GA has successfully been used to solve some problems such as fitting nonlinear regression to data, optimizing simulation models, solving systems of nonlinear equations, and machine learning (Deb 1998). Generally, a GA has five major components to solve a particular problem that are briefly described below:

  • 1

    At the first, n chromosomes generate a population randomly that are known as candidate solutions to the problem.

  • 2
    A special fitness function evaluates the fitness of each chromosome. In the present study, efficiency coefficient (EC) was used as the fitness function and it can be written as:
    formula
    (13)
    where N represents the total number of a testing data and is the predicted value. is the observed value and is the mean of the observed values.
  • 3

    The following steps will be repeated until n offsprings have been created:

    • (a)

      Selection: This operator selects the best chromosomes in pairs from the population to play the role of parents and reproduce two offspring. The more appropriate chromosomes have more chances to be selected.

    • (b)

      Crossover: This operator randomly chooses a locus between a couple of chromosomes to form two offspring.

    • (c)

      Mutation: This operator creates new chromosomes by flipping some of the bits in the chromosomes randomly.

  • 4

    Replace the current population with the new population.

  • 5

    If the stopping condition is satisfied, the best solution is returned in the current population, otherwise step 2 should be performed again.

The applied GA method settings in the present study are shown in Table 3.

Table 3

Genetic algorithm settings

Population size 250 
Number of generations 10 
Elitism 12 
Crossover probability 0.8 
Mutation probability 0.1 
Crossover function Scatter 
Mutation function Gaussian 
Population size 250 
Number of generations 10 
Elitism 12 
Crossover probability 0.8 
Mutation probability 0.1 
Crossover function Scatter 
Mutation function Gaussian 

Genetic algorithm-based support vector machine

In this study, at first, the training data (input and target parameters) are presented to the GA-SVM algorithm. Then, GA randomly generates an initial population of unknown SVM's parameters (, , and ) to determine their optimal values to approach the best prediction with the lowest error and the highest accuracy. The fitness function examines the performance of each model. The secondary population of SVM's parameters is created by using the operators of GA (mutation, crossover, and selection) to obtain the optimal values of parameters and then these parameters are introduced to the SVM algorithm, again. This cycle is continued until the value of the fitness function is near or equal to the stopping conditions of the algorithm. Therefore, model outputs are expected to be closer to the target values ​​at each cycle. In the GA-SVM algorithm, both algorithms operate separately but help each other in order to simplify the problem. In other words, first, SVM starts modeling by using the random parameters generated by GA, and GA continues the procedure of modeling until the optimal values of SVM's parameters are obtained. In this method, the GA algorithm tries to estimate the optimal combination of three parameters (, and ) in each cycle. C is known as a regularization parameter that must control the trade-off between maximizing the margin and minimizing the training error. Low C values will place insufficient stress on fitting the training data and high values of C make the algorithm over-fit the training data (Noori et al. 2011). Nevertheless, according to Wang et al. (2003), it can be concluded that the prediction error is rarely influenced by C. denotes the optimal width of the kernel function, while RBF with large allows the support vector to have a strong impact over a larger area. The type of noise present in data determines the optimal value for , which is usually unknown. There is a practical consideration of the number of resulting support vectors, even if enough knowledge of the noise is available for selecting an optimal value for (Liu et al. 2006). In the GA-SVM hybrid algorithm, GA automatically starts finding the mentioned parameters of SVM and provides the optimal values, while determining the optimal values of parameters in the SVM algorithm was done by trial-and-error process. The cross-validation, which is an improved version of the grid search method, described by Hsu et al. (2010), was used to find these three parameters. In ν-fold cross-validation, after the training set was divided into ν subsets of equal size, one subset is tested sequentially by applying the classifier trained on the remaining ν − 1 subset. Therefore, each instance of the whole training set is estimated once so the cross-validation accuracy is the percentage of correctly classified data. The general flowchart of GA-SVM is illustrated in Figure 4.

Figure 4

General flowchart of GA-SVM algorithm.

Figure 4

General flowchart of GA-SVM algorithm.

Close modal

In the present study, SVM and GA-SVM were applied by using RBF kernel function and input variables. Table 2 shows that all parameters used in this study have a right-skewed distribution. On the other hand, according to Figure 5, there is an abundance of outliers in the target and input parameters except and . Those observations which are uncommon and do not conform to the pattern of the majority of the data are called outliers (Rousseeuw & Van Zomeren 1990). The existence of outliers can cause increased error rates and reduce the accuracy of prediction. It can also lead to considerable distortions of statistic estimates when using either parametric or nonparametric tests (Zimmerman 1994, 1995, 1998). One of the simplest methods to tackle this problem is logarithmic transformations of parameters individually or collectively (Hubert & Van der Veeken 2008). Therefore, to reduce the negative effects of skewness and outliers on modeling, the whole dataset had been transformed into logarithmic scale and then the logarithmic parameters were used for modeling.

Figure 5

Boxplots of all parameters with outliers (*).

Figure 5

Boxplots of all parameters with outliers (*).

Close modal

Model evaluation

In this study, both SVM and GA-SVM were used to estimate the TMC. The performances of these two models are assessed by evaluating the scatter plots between the observed and predicted results. In addition, the discrepancy ratio (DR), the root mean square error (RMSE), the mean of the absolute error (ME) and the accuracy were used as statistical parameters to evaluate the performance of SVM, GA-SVM, and empirical models. Statistical indexes that were used in this study are expressed as:
formula
(14)
formula
(15)
formula
(16)
where and are predicted and observed TMCs, respectively, and N is the total number of data points. If DR is equal to zero, there will be an exact match between the observed and predicted values. An overestimation () or underestimation () otherwise occurs. Previous researchers reported the percentage of DR values between −0.3 and 0.3 as an accuracy index (Seo & Cheong 1998; Kashefipour & Falconer 2002). In this research, in order to better evaluate the models’ performance and accuracy, percentages of DR values between −0.15 and 0.15 were used as an accuracy index (Figure 6). As well, DR < −0.15 and DR > 0.15 have been considered as underestimation and overestimation beyond the precision range, respectively. A comparison of DR frequency could be used to determine the symmetry and skewness of TMC estimation by different models.
Figure 6

Comparison of accuracy index between previous studies and the current study.

Figure 6

Comparison of accuracy index between previous studies and the current study.

Close modal

For estimating TMC by using SVM, as was mentioned before, we first need to find the optimal values of three adjustable parameters of SVM (, and ). During the grid search, all combinations of were tested for each cross-validation routine, where these parameters all ranged from 0 to 120. Finally, the optimum values of these three parameters were determined by using both GA and grid search algorithms. These values are presented in Table 4. According to Table 4, although both GA and grid search algorithms estimate parameter C to be approximately the same, their estimations were different for the other two parameters. It should be noted that GA does not estimate the optimal value of each parameter separately. This algorithm estimates only the optimal combination of the three parameters.

Table 4

Optimal parameters of GA-SVM and SVM models

ModelsMethod
GA-SVM GA 3.01 0.15 0.47 
SVM Grid Search 3.00 0.01 1.00 
ModelsMethod
GA-SVM GA 3.01 0.15 0.47 
SVM Grid Search 3.00 0.01 1.00 

The performances of SVM, GA-SVM, and the previous methods in TMC estimation by using the mentioned statistical indexes are presented in Table 5.

Table 5

Performances of various methods on TMC estimation

Models(DR < −0.15)(−0.15 < DR < 0)(0 < DR < 0.15)(0.15 < DR)Accuracy%MAERMSE
Fischer & Park (1967)  15.086 9.052 19.397 56.466 28.448 0.228 0.270 
Yotsukura et al. (1970)  2.155 1.724 6.466 89.655 8.190 0.588 0.626 
Chau (2000)  19.397 12.931 52.586 15.086 65.517 0.180 0.255 
Ahmad (2007)  25.431 28.017 41.810 4.741 69.828 0.169 0.273 
Jeon et al. (2007)  12.931 13.362 31.034 42.672 44.397 0.188 0.233 
Azamathulla & Ahmad (2012)  31.034 31.466 35.345 2.155 66.810 0.180 0.287 
Aghababaei et al. (2017)  12.069 37.931 42.672 7.328 80.603 0.096 0.148 
Zahiri & Nezaratian (2020)  11.638 31.466 44.397 12.500 75.862 0.113 0.149 
GA-SVM (Train) 5.747 42.529 50.000 1.724 92.529 0.066 0.107 
GA-SVM (Test) 10.345 32.759 50.000 6.897 82.759 0.097 0.139 
SVM (Train) 5.747 42.529 48.851 2.874 91.379 0.044 0.096 
SVM (Test) 12.069 32.759 48.276 6.897 81.034 0.097 0.152 
Models(DR < −0.15)(−0.15 < DR < 0)(0 < DR < 0.15)(0.15 < DR)Accuracy%MAERMSE
Fischer & Park (1967)  15.086 9.052 19.397 56.466 28.448 0.228 0.270 
Yotsukura et al. (1970)  2.155 1.724 6.466 89.655 8.190 0.588 0.626 
Chau (2000)  19.397 12.931 52.586 15.086 65.517 0.180 0.255 
Ahmad (2007)  25.431 28.017 41.810 4.741 69.828 0.169 0.273 
Jeon et al. (2007)  12.931 13.362 31.034 42.672 44.397 0.188 0.233 
Azamathulla & Ahmad (2012)  31.034 31.466 35.345 2.155 66.810 0.180 0.287 
Aghababaei et al. (2017)  12.069 37.931 42.672 7.328 80.603 0.096 0.148 
Zahiri & Nezaratian (2020)  11.638 31.466 44.397 12.500 75.862 0.113 0.149 
GA-SVM (Train) 5.747 42.529 50.000 1.724 92.529 0.066 0.107 
GA-SVM (Test) 10.345 32.759 50.000 6.897 82.759 0.097 0.139 
SVM (Train) 5.747 42.529 48.851 2.874 91.379 0.044 0.096 
SVM (Test) 12.069 32.759 48.276 6.897 81.034 0.097 0.152 

Along with MAE, RMSE, and accuracy indexes, the balance between overestimation and underestimation values is also another important point in analyzing the models' performances. According to Table 5, among the previous regression models, the two models of Yotsukura et al. (1970) and Fischer & Park (1967), had the lowest performances in estimating the TMC with the accuracy of 8% and 28.5%, respectively. The two models of Aghababaei et al. (2017) and Zahiri & Nezaratian (2020) were able to have accurate performances in estimating TMC. The model of Aghababaei et al. (2017), based on GPSR method, with an accuracy of 80% and RMSE and MAE values of 0.148 and 0.096, respectively, and the simple data-driven-based model proposed by Zahiri & Nezaratian (2020) with a relatively good accuracy (75.8%) and the balance between overestimation and underestimation values were the most accurate regression-based models available to estimate this coefficient. Both GA-SVM and SVM algorithms had genuinely accurate and relatively similar performances. In the testing stage, both of them had the least error rates and the highest accuracy compared to the previous regression-based models. It should also be noted that although both models were based on the SVM algorithm, GA-SVM compared to SVM was able to improve the accuracy of the TMC estimation gently, in both training and testing stages by 1.15% and 1.7%, respectively. On the other hand, the grid search method is more time-consuming than GA, which make the GA-SVM model chosen for estimating TMC in this study. A comparison of the DR values of all expressions along with developed SVM and GA-SVM models is demonstrated in Figure 7. In addition, Figure 8 shows the performances of the developed SVM and GA-SVM in estimating the TMC for the two training and testing stages.

Figure 7

Comparison of the DR values of different methods.

Figure 7

Comparison of the DR values of different methods.

Close modal
Figure 8

The observed and predicted TMC (m2/s) values by: (a) SVM in the training stage, (b) SVM in the testing stage, (c) GA-SVM in the training stage, and (d) GA-SVM in the testing stage.

Figure 8

The observed and predicted TMC (m2/s) values by: (a) SVM in the training stage, (b) SVM in the testing stage, (c) GA-SVM in the training stage, and (d) GA-SVM in the testing stage.

Close modal

Based on Figure 7, the superiority of GA-SVM and SVM performance is obvious and both models have lower overestimation and underestimation values than the models of Aghababaei et al. (2017) and Zahiri & Nezaratian (2020). In addition, in Figure 8, the estimating accuracy by SVM and GA-SVM models are shown in training and testing stages, separately. The dataset used in this study included characteristics of straight and meandering streams. According to Table 6, the performance of both SVM and GA-SVM models in both straight and meandering streams was more accurate than the regression-based models. All models performed better in estimating the TMC in straight streams than meandering ones.

Table 6

Performances of various models using data of straight and meandering streams

ModelsStraight
Meandering
Accuracy%MAERMSEAccuracy%MAERMSE
Aghababaei et al. (2017)  85.246 0.082 0.124 63.265 0.150 0.216 
Zahiri & Nezaratian (2020)  86.339 0.089 0.115 36.735 0.200 0.235 
GA-SVM 93.443 0.063 0.099 77.551 0.113 0.164 
SVM 91.803 0.049 0.098 77.551 0.083 0.155 
ModelsStraight
Meandering
Accuracy%MAERMSEAccuracy%MAERMSE
Aghababaei et al. (2017)  85.246 0.082 0.124 63.265 0.150 0.216 
Zahiri & Nezaratian (2020)  86.339 0.089 0.115 36.735 0.200 0.235 
GA-SVM 93.443 0.063 0.099 77.551 0.113 0.164 
SVM 91.803 0.049 0.098 77.551 0.083 0.155 

Sensitivity analysis

Sensitivity analysis helps researchers to determine which parameter has the most effect on reducing output uncertainty, and/or which parameters are negligible and can be eliminated from the final model (Nezaratian et al. 2018). In this study, a sensitivity analysis method was applied to determine the effect of each parameter on the performance of GA-SVM as the most accurate model in the TMC estimation. Five scenarios of the input parameter combinations were introduced to the GA-SVM algorithm for the TMC estimation. Table 7 presents the combination of inputs, absent parameters, SVM parameters, and the performance of each scenario in the testing stage, respectively.

Table 7

Sensitivity analysis of GA-SVM scenarios

ScenarioInputsAbsentParameters (Accuracy%MAERMSEΔ;Accuracy%
U/U*, Fr, Sn W/H 7.75, 0.11, 0.30 84.483 0.064 0.110 1.725 
W/H, Fr, Sn U/U* 5.47, 0.27, 0.20 86.207 0.074 0.131 3.448 
W/H, U/U*, Sn Fr 4.38, 0.19, 0.25 89.655 0.062 0.117 6.896 
W/H, U/U*, Fr Sn 2.33, 0.33, 1.59 81.034 0.087 0.124 −1.725 
W/H, Sn U/U*, Fr 3.50, 0.47, 0.67 91.379 0.071 0.137 8.620 
ScenarioInputsAbsentParameters (Accuracy%MAERMSEΔ;Accuracy%
U/U*, Fr, Sn W/H 7.75, 0.11, 0.30 84.483 0.064 0.110 1.725 
W/H, Fr, Sn U/U* 5.47, 0.27, 0.20 86.207 0.074 0.131 3.448 
W/H, U/U*, Sn Fr 4.38, 0.19, 0.25 89.655 0.062 0.117 6.896 
W/H, U/U*, Fr Sn 2.33, 0.33, 1.59 81.034 0.087 0.124 −1.725 
W/H, Sn U/U*, Fr 3.50, 0.47, 0.67 91.379 0.071 0.137 8.620 

As presented in Table 7, the effect of eliminating each input parameter on accuracy of final GA-SVM model was determined. In the table above, ΔAccuracy% expresses the difference between the final accuracy of each scenario and the overall accuracy in the testing stage. It should be noticed that the above method significantly depends on the mathematical and theoretical structure of GA-SVM and may not be able to introduce the most effective parameter on TMC. However, analyzing Table 7 could help us, to some extent, on the effect of each input parameter on TMC estimation. The logic of input combination in scenario 5 was based on Figure 2. According to this figure, and have the highest correlation with the dimensionless parameter of the TMC while the lowest correlation belongs to and , respectively. Therefore, scenario 5 was used to measure the impact of removing the least correlated parameters on modeling TMC by GA-SVM. According to Table 7, in scenario 1, by eliminating from the input parameters, the accuracy increases by 1.725%. However, in scenario 2, when was replaced with in the input variables, the accuracy was improved by 3.488%. In addition, using the same analysis and considering scenario 3, it can be deduced that is the least effective parameter on TMC estimation by using the GA-SVM algorithm. According to scenario 4, it can also be concluded that is a most efficient parameter in the process of modeling TMC. In scenario 5, only inputs which had a correlation coefficient above 0 were used, so and were eliminated from the process. The result showed that there was a significant improvement in the accuracy of the final model, which increase the modeling accuracy by 8.26%. Table 7 demonstrated that reducing the number of input variables with low correlation with the target improved the performance of the final GA-SVM model. Eliminating the low correlated input variables could decrease the complexity of the modeling process and increase the accuracy. This finding agreed with the results of Zahiri & Nezaratian (2020) and Jeon et al. (2007), which showed that and are the most influential parameters in estimating the TMC, respectively.

In this study, SVM and GA-SVM algorithms were developed to estimate the transverse mixing coefficient that plays an important role in modeling the pollutant release into streams. For this purpose, three statistical indexes (accuracy, RMSE, and MAE) were used to determine the performance of different models. The results showed the superiority of the proposed model compared to well-known regression-based models. The results also showed that the two models proposed by Aghababaei et al. (2017) and Zahiri & Nezaratian (2020) had the highest accuracy in estimating the TMC, respectively. Dividing the dataset into two groups (straight and meandering streams) showed that SVM and GA-SVM are still more reliable than the previous models. In this study, the grid search method was used to develop the SVM algorithm and was much more time-consuming than the GA algorithm. Therefore, the GA-SVM model was chosen as the best model to estimate the TMC in streams. Then, a sensitivity analysis was performed to determine the most effective input parameters in estimating the TMC by GA-SVM. Based on the sensitivity analysis, and had the least impact on GA-SVM performance in estimating TMC, and eliminating these two parameters improved the accuracy of the TMC estimation.

All relevant data are available from an online repository or repositories (https://data.mendeley.com/datasets/2mm7jmp2g5/1).

Aghababaei
M.
Etemad-Shahidi
A.
Jabbari
E.
Taghipour
M.
2017
Estimation of transverse mixing coefficient in straight and meandering streams
.
Water Resources Management
31
(
12
),
3809
3827
.
Ahmad
Z.
2007
Two-dimensional Mixing of Pollutants in Open Channels
.
A technical report submitted to DST
,
New Delhi
,
India
.
Ahmad
Z.
2008
Finite volume model for steady-state transverse mixing in streams
.
Journal of Hydraulic Research
46
(
suppl. 1
),
72
80
.
Ahmad
Z.
Azamathulla
H. M.
Zakaria
N. A.
2011
ANFIS-based approach for the estimation of transverse mixing coefficient
.
Water Science and Technology
63
(
5
),
1004
1009
.
Alizadeh
M. J.
Ahmadyar
D.
Afghantoloee
A.
2017
Improvement on the existing equations for predicting longitudinal dispersion coefficient
.
Water Resources Management
31
(
6
),
1777
1794
.
Antonopoulos
V. Z.
Georgiou
P. E.
Antonopoulos
Z. V.
2015
Dispersion coefficient prediction using empirical models and ANNs
.
Environmental Processes
2
(
2
),
379
394
.
Azamathulla
H. M.
Ahmad
Z.
2012
Gene-expression programming for transverse mixing coefficient
.
Journal of Hydrology
434
,
142
148
.
Azamathulla
H.
Ghani
A.
2011
Genetic programming for predicting longitudinal dispersion coefficients in streams
.
Water Resources Management
25
(
6
),
1537
1544
.
Azamathulla
H. M.
Wu
F. C.
2011
Support vector machine approach for longitudinal dispersion coefficients in natural streams
.
Applied Soft Computing
11
(
2
),
2902
2905
.
Baek
K. O.
Seo
I. W.
2013
Empirical equation for transverse dispersion coefficient based on theoretical background in river bends
.
Environmental Fluid Mechanics
13
(
5
),
465
477
.
Beltaos
S.
1979
Transverse mixing in natural streams
.
Canadian Journal of Civil Engineering
6
(
4
),
575
591
.
Beltaos
S.
1980
Transverse mixing tests in natural streams
.
Journal of the Hydraulics Division
106
(
10
),
1607
1625
.
Boxall
J. B.
Guymer
I.
2003
Analysis and prediction of transverse mixing coefficients in natural channels
.
Journal of Hydraulic Engineering
129
(
2
),
129
139
.
Chau
K. W.
2000
Transverse mixing coefficient measurements in an open rectangular channel
.
Advances in Environmental Research
4
(
4
),
287
294
.
Deb
K
.
1998
Genetic algorithm in search and optimization: the technique and applications
. In
Proceedings of International Workshop on Soft Computing and Intelligent Systems
.
Machine Intelligence Unit, Indian Statistical Institute Calcutta
,
India
, pp.
58
87
.
Demetracopoulos
A. C.
1994
Computation of transverse mixing in streams
.
Journal of Environmental Engineering
120
(
3
),
699
706
.
Deng
Z. Q.
Singh
V. P.
Bengtsson
L.
2001
Longitudinal dispersion coefficient in straight rivers
.
Journal of Hydraulic Engineering
127
(
11
),
919
927
.
Etemad-Shahidi
A.
Taghipour
M.
2012
Predicting longitudinal dispersion coefficient in natural streams using M5′ model tree
.
Journal of Hydraulic Engineering
138
(
6
),
542
554
.
Fischer
H. B.
1967
The mechanics of dispersion in natural streams
.
Journal of the Hydraulics Division
93
(
6
),
187
216
.
Fischer
H. B.
Park
M.
1967
Transverse Mixing in A Sand-bed Channel
.
US Geological Survey
Professional Paper, 267–272
.
Fischer
H. B.
List
J. E.
Koh
C. R.
Imberger
J.
Brooks
N. H.
1979
Mixing in Inland and Coastal Waters
.
Academic Press
,
New York
.
Goldberg
D. E.
1989
Genetic Algorithms in Search, Optimization and Machine Learning
.
Addison-Wesley
,
New York
,
USA
.
Haghiabi
A. H.
Nasrolahi
A. H.
Parsaie
A.
2018
Water quality prediction using machine learning methods
.
Water Quality Research Journal
53
(
1
),
3
13
.
Holland
J. H.
1975
Adaptation in Natural and Artificial Systems
.
University of Michigan Press
,
Ann Arbor, MI
,
USA
.
Holley
E. R.
Abraham
G.
1973
Laboratory studies on transverse mixing in rivers
.
Journal of Hydraulic Research
11
(
3
),
219
253
.
Hsu
C. C.
Chen
M. C.
Chen
L. S.
2010
Intelligent ICA–SVM fault detector for non-Gaussian multivariate process monitoring
.
Expert Systems with Applications
37
(
4
),
3264
3273
.
Huai
W.
Shi
H.
Yang
Z.
Zeng
Y.
2018
Estimating the transverse mixing coefficient in laboratory flumes and natural rivers
.
Water, Air, & Soil Pollution
229
(
8
),
252
.
Hubert
M.
Van der Veeken
S.
2008
Outlier detection for skewed data
.
Journal of Chemometrics: A Journal of the Chemometrics Society
22
(
3–4
),
235
246
.
Jeon
T. M.
Baek
K. O.
Seo
I. W.
2007
Development of an empirical equation for the transverse dispersion coefficient in natural streams
.
Environmental Fluid Mechanics
7
(
4
),
317
329
.
Kashefipour
S. M.
Falconer
R. A.
2002
Longitudinal dispersion coefficients in natural channels
.
Water Research
36
(
6
),
1596
1608
.
Krishnappan
B. G.
Lau
Y. L.
1977
Transverse mixing in meandering channels with varying bottom topography
.
Journal of Hydraulic Research
15
(
4
),
351
370
.
Lau
Y. L.
Krishnappan
B. G.
1981
Modeling transverse mixing in natural streams
.
Journal of the Hydraulics Division
107
(
2
),
209
226
.
Lee
M. E.
Seo
I. W.
2013
Spatially variable dispersion coefficients in meandering channels
.
Journal of Hydraulic Engineering
139
(
2
),
141
153
.
Li
X. Z.
Kong
J. M.
2014
Application of GA-SVM method with parameter optimization for landslide development prediction
.
Natural Hazards and Earth System Sciences
14
(
3
),
525
.
Li
X.
Liu
H.
Yin
M.
2013
Differential evolution for prediction of longitudinal dispersion coefficients in natural streams
.
Water Resources Management
27
(
15
),
5245
5260
.
Liu
H. B.
Jiao
Y. B.
2011
Application of genetic algorithm-support vector machine (GA-SVM) for damage identification of bridge
.
International Journal of Computational Intelligence and Applications
10
(
4
),
383
397
.
Nezaratian
H.
Zahiri
J.
Kashefipour
S. M.
2018
Sensitivity analysis of empirical and data-driven models on longitudinal dispersion coefficient in streams
.
Environmental Processes
5
(
4
),
833
858
.
Noori
R.
Karbassi
A. R.
Moghaddamnia
A.
Han
D.
Zokaei-Ashtiani
M. H.
Farokhnia
A.
Gousheh
M. G.
2011
Assessment of input variables determination on the SVM model performance using PCA, gamma test, and forward selection techniques for monthly stream flow prediction
.
Journal of Hydrology
401
(
3–4
),
177
189
.
Parsaie
A.
Haghiabi
A. H.
2015
Predicting the longitudinal dispersion coefficient by radial basis function neural network
.
Modeling Earth Systems and Environment
1
(
4
),
1
8
.
Parsaie
A.
Haghiabi
A. H.
2017a
Mathematical expression of discharge capacity of compound open channels using MARS technique
.
Journal of Earth System Science
126
(
2
),
20
.
Parsaie
A.
Haghiabi
A. H.
2017b
Numerical routing of tracer concentrations in rivers with stagnant zones
.
Water Science and Technology: Water Supply
17
(
3
),
825
834
.
Parsaie
A.
Emamgholizadeh
S.
Azamathulla
H. M.
Haghiabi
A. H.
2018
ANFIS-based PCA to predict the longitudinal dispersion coefficient in rivers
.
International Journal of Hydrology Science and Technology
8
(
4
),
410
424
.
Parsaie
A.
Haghiabi
A. H.
Moradinejad
A.
2019
Prediction of scour depth below river pipeline using support vector machine
.
KSCE Journal of Civil Engineering
23
(
6
),
2503
2513
.
Pourbasheer
E.
Riahi
S.
Ganjali
M. R.
Norouzi
P.
2009
Application of genetic algorithm-support vector machine (GA-SVM) for prediction of BK-channels activity
.
European Journal of Medicinal Chemistry
44
(
12
),
5023
5028
.
Riahi-Madvar
H.
Ayyoubzadeh
S. A.
Khadangi
E.
Ebadzadeh
M. M.
2009
An expert system for predicting longitudinal dispersion coefficient in natural streams by using ANFIS
.
Expert Systems with Applications
36
(
4
),
8589
8596
.
Roushangar
K.
Koosheh
A.
2015
Evaluation of GA-SVR method for modeling bed load transport in gravel-bed rivers
.
Journal of Hydrology
527
,
1142
1152
.
Rousseeuw
P. J.
Van Zomeren
B. C.
1990
Unmasking multivariate outliers and leverage points
.
Journal of the American Statistical Association
85
(
411
),
633
639
.
Rutherford
J. C.
1994
Longitudinal Dispersion. River Mixing
.
John Wiley and Sons
,
Chichester
,
UK
.
Sattar
A. M.
Gharabaghi
B.
2015
Gene expression models for prediction of longitudinal dispersion coefficient in streams
.
Journal of Hydrology
524
,
587
596
.
Seo
I. W.
Cheong
T. S.
1998
Predicting longitudinal dispersion coefficient in natural streams
.
Journal of Hydraulic Engineering
124
(
1
),
25
32
.
Sharma
H.
Ahmad
Z.
2014
Transverse mixing of pollutants in streams: a review
.
Canadian Journal of Civil Engineering
41
(
5
),
472
482
.
Stefanovic
D. L.
Stefan
H. G.
2001
Accurate two-dimensional simulation of advective-diffusive-reactive transport
.
Journal of Hydraulic Engineering
127
(
9
),
728
737
.
Vapnik
V. N.
1995
The Nature of Statistical Learning Theory
.
Springer-Verlag
,
New York
,
USA
.
Wang
W.
Xu
Z.
Lu
W.
Zhang
X.
2003
Determination of the spread parameter in the Gaussian kernel for classification and regression
.
Neurocomputing
55
(
3–4
),
643
663
.
Wang
W. C.
Xu
D. M.
Chau
K. W.
Chen
S.
2013
Improved annual rainfall-runoff forecasting using PSO–SVM model based on EEMD
.
Journal of Hydroinformatics
15
(
4
),
1377
1390
.
Yotsukura
N.
Sayre
W. W.
1976
Transverse mixing in natural channels
.
Water Resources Research
12
(
4
),
695
704
.
Yotsukura
N.
Fischer
H. B.
Sayre
W. W.
1970
Measurement of Mixing Characteristics of the Missouri River Between Sioux City, Iowa, and Plattsmouth, Nebraska
.
Water Supply Paper No. 1899-G
.
USGPO
.
Zahiri
J.
Nezaratian
H.
2020
Estimation of transverse mixing coefficient in streams using M5, MARS, GA, and PSO approaches
.
Environmental Science and Pollution Research
27
,
14553
14566
.
Zimmerman
D. W.
1994
A note on the influence of outliers on parametric and nonparametric tests
.
The Journal of General Psychology
121
(
4
),
391
401
.
Zimmerman
D. W.
1995
Increasing the power of nonparametric tests by detecting and downweighting outliers
.
The Journal of Experimental Education
64
(
1
),
71
78
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).

Supplementary data